Structure of amorphous iron-based coatings processed by HVOF

Materials Chemistry and Physics 85 (2004) 113–119
Structure of amorphous iron-based coatings processed by
HVOF and APS thermally spraying
M. Cherigui, H.I. Feraoun, N.E. Feninehe∗ , H. Aourag, C. Coddet
LERMPS, UTBM, 90010 Belfort Cedex, France
Received 14 July 2003; received in revised form 7 December 2003; accepted 17 December 2003
Abstract
High-velocity oxy-fuel (HVOF) and air-plasma spraying (APS) were used to produce Fe3 Si and FeNb alloy coatings. Adjusting the
spraying conditions, amorphous coatings were obtained from the FeNb powder. However, for the Fe3 Si powder, crystalline deposits were
produced in all cases. Boron additives were used to improve the aptitude of the Fe3 Si alloy to form an amorphous phase. In this perspective,
first-principle calculations were elaborated to investigate the electronic structure of crystalline FeNb and Fe3 Si. It was shown that the
introduction of boron impurities into the alloy matrix lead to the lowering of the structural stability, and made its electronic density of state
more comparable to the corresponding FeNb structure.
© 2004 Elsevier B.V. All rights reserved.
Keywords: Amorphous coatings; Thermal spraying; Magnetic properties; Fe3 Si; FeNb; Rapid quenching; Intermetallics; Electronic structures
1. Introduction
The development of amorphous and nanocrystalline materials has attracted significant interest in the field of new
materials design. Indeed, the magnetic, chemical and mechanical properties of materials are largely enhanced when
the size of crystallites becomes nanometric. In addition,
the absence of crystal structure implicates a macroscopic
behavior fairly different from that of corresponding to the
polycrystalline state, especially, mechanical and magnetic
properties.
Amorphous metallic materials can be obtained by rapid
quenching [1] in order to prevent traditional solidification
phenomena from occurring. Thus, one can obtain materials with a disordered structure beyond some interatomic
distances. As a consequence, the microstructure does not
contain grains as in polycrystals, and so no grain boundaries, elements that control to a large extent the macroscopic
behavior. Accordingly, some amorphous metallic alloys exhibit exceptional mechanical, chemical and magnetic properties due to the lack of long-range order in their atomic
arrangement [2]. However, industrial applications of these
amorphous alloys have been restricted due to the difficulties
encountered in the production of bulk quantities and the
∗ Corresponding author. Tel.: +33-3-84-58-31-16;
fax: +33-3-84-58-32-86.
E-mail address: [email protected] (N.E. Feninehe).
0254-0584/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.matchemphys.2003.12.017
limitations in thickness caused by rapid quenching methods
such as melt spinning [2]. There exists a large number of
techniques to produce amorphous alloys other than rapid
quenching, namely vapor condensation on cold substrate,
cathode sputtering, ion bombardments, electrolysis, and
amorphization methods in solid state [3]. Thermal spraying
has become a very advantageous method owing to the high
deposition rates and low operation costs involved [4,5]. The
thermal spraying process permits the use of complex geometry substrates. Then its use is extended to a wide range of applications. Furthermore, free standing near net shape forms
can also be produced by this droplet consolidation technique
[6,7].
The interest found in Fe3 Si alloys was derived by the
intriguing magnetic properties of this compound, specifically its magnetic susceptibility χ(T) which is small at low
temperatures T, but grows rapidly until it peaks near 500 K
and starts decreasing slowly above this value following the
Curie–Weiss law. Benoit [8] proposed that the broad maximum susceptibility is due to an antiferromagnetic transition,
but the results of subsequent neutron scattering [9] show
that the lack of crystalline structure plays a key role in this
special magnetic behavior.
In order to take advantage of these properties, amorphous
Fe3 Si coatings were produced using thermal spraying.
Subsequent trial tests with different spraying conditions
were not successful for this alloy, whereas it has been easy
to obtain amorphous FeNb coatings. Following several
114
M. Cherigui et al. / Materials Chemistry and Physics 85 (2004) 113–119
Table 1
Spraying parameters
Parameters
(l min−1 )
Oxygen flow
Methane flow (l min−1 )
Carrying N2 (l min−1 )
Feed rate (g min−1 )
Spray distance (mm)
Plate speed (m s−1 )
Tube
Plate
420
140
20
35
300
420
140
20
35
300
≈1
suggestions [1], some additive elements within the Fe3 Si
matrix can lead to an easier amorphization. Looking into
the literature, it was found that Fe3 Si–B amorphous ribbons
have been widely used as a magnetic core material and
there are also many reports on their crystallization behavior
[10,11]. Consequently, a structural and electronic study of
crystalline Fe3−x SiBx and FeNb alloys were initiated via
first-principle total-energy simulation method. The suppercell approach was used to examine the local effect of boron
atoms in the Fe3 Si structure, and compare the appropriate
Fig. 1. Morphology and structure of FeNb coatings. (a) SEM image and EDS analysis of the white zone (b), the non-molten particles (c) and the gray
zone (d).
M. Cherigui et al. / Materials Chemistry and Physics 85 (2004) 113–119
115
properties to those of FeNb alloy. In this study, a FeNb
alloy is used as a reference material since its amorphization
was relatively easy as well as it was stated elsewhere [12].
2. Materials and experiments
High-velocity oxy-fuel (HVOF) and air-plasma spraying
(APS) were used to spray Fe–50 wt.% Nb (0–44 ␮m) and
Fe–6.5 wt.% Si (0–150 ␮m) particle powders. The copper
substrate used is a 70 mm × 25 mm and 0.8 mm thick sheet
fixed on a port sample moved at a velocity of 200 rpm. The
spraying experimental conditions are given in Table 1. To
avoid significant heating during the progress of the process,
the substrate was cooled on its backside using compressed
air. Several passes (15–25) of the torch were necessary to
produce 200–300 ␮m thick coatings.
The as-sprayed coatings were cut and polished using standard procedures for thermal sprayed metal coatings. Optical
microscopy was used to obtain low magnification images
using a Nikon Epiphot 200 microscope. Scanning electron
microscopy (SEM) was used to obtain clear images. The
procedure adopted for image acquisition entailed a backscattered electron imaging mode for higher contrast between
pores, oxide areas and metal matrix. Energy-dispersive spectrometry (EDS) analysis was employed while imaging on
SEM to obtain the elemental composition at different areas
of the coating cross-section. Starting powders were also investigated using X-ray diffraction (XRD). It was conducted
using a PHILIPS X’PERT diffractometer at a scanning rate
of 3◦ min−1 and with copper radiation (Cu K␣).
3. Experimental results
Figs. 1a and 2a are SEM images of FeNb and Fe3 Si
coatings showing the typical aspect of the as-sprayed structure. The deposits were fairly dense and well-bonded on the
copper substrate. The individual splats were clearly visible
within the coatings. For a given coating, the morphology of
these splats is well different from one area to the other, and
this is true for both types of coatings. Some of them have
melted and flowed extensively on impact whereas, for others, the particulate nature of the starting material was still
clearly apparent.
In metallic coatings, it is difficult to identify whether
the dark areas observed in the microstructure represent gas
porosity, interlamellar pores or oxides. EDS analysis can be
used to obtain the elemental composition from different areas. This analysis made it possible to detect the presence of
two phases for the FeNb coating, one in white color (36.3%
Fe and 63.2% Nb) and other in gray color (20.2% Fe and
78.8% Nb). For Fe3 Si coatings, the presence of Si with almost the same percentage as in the powder (78.3% Fe and
5.9% Si) can be noted. The presence of other elements like
S, Cr, Pd and C in poor amounts was also observed.
Fig. 2. Morphology and structure of Fe3 Si coatings. (a) SEM image and
EDS analysis of the gray zone (b) and the white zone (c).
Image analysis was applied to these coatings produced by
the APS process in order to quantify the content of porosity
and oxides. For the Fe3 Si coatings, Fig. 2 shows a less dense
deposit than the FeNb one. The porosity of the Fe3 Si coating
is superior than that of FeNb: we have about 15 ± 2.5 and
9.5 ± 2% of porosity ratios, respectively. The presence of
these various zones in the FeNb coatings suggests that the
structure is completely different from a simple crystalline
structure and XRD results confirmed these conclusions.
Fig. 3 illustrates XRD patterns for FeNb alloy. Regarding
to the peak intensities and their broadening, we noticed a
mixture of amorphous and extremely fine crystalline phase
(nanocrystalline). This partially amorphous structure of
FeNb type with the presence of NbO2 and Nb2 O5 oxides
results from the use of the HVOF technique where the
oxidation phenomena are significant. However, for Fe3 Si
alloy, the X-ray characterization shows that the structure is
116
M. Cherigui et al. / Materials Chemistry and Physics 85 (2004) 113–119
Fig. 3. X-ray diffraction patterns for (a) FeNb powder, (b) HVOF coatings
and (c) APS coatings.
perfectly crystalline: the basically cubic structure ␣1 –Fe3 Si
(DO3 ), which is characteristic for the low-silicon content
Fe3 Si alloys is predominant (Fig. 4). These results agree
with those obtained by rapidly quenched Fe100−x Six alloy
[13,14] and mechanically alloyed Fe3 Si [15].
Fig. 5 shows standard hysteresis loops for the Fe3 Si coatings. We can note that this Fe3 Si is a soft ferromagnet with
an average coercivity Hc of 4, 6.7 and 8 Oe for powder,
Fig. 5. Hysteresis loops for (a) Fe3 Si powder, (b) HVOF coatings and (c)
APS coatings.
APS and HVOF coatings, respectively. On the other hand,
the magnetic characterization shows the super paramagnetic
state of the FeNb alloy coating.
4. Calculation details
4.1. Method
Fig. 4. X-ray diffraction patterns for (a) Fe3 Si powder, (b) HVOF coatings
and (c) APS coatings.
Our calculations were carried using the full potential linear augmented plane wave (FPLAPW), a first-principle total
energy density functional theory (DFT)-based method. We
used the scalar relativistic version without spin orbit coupling [16] as embodied in the WIEN2k code [17] which
is thought to be from the most accurate implementations
M. Cherigui et al. / Materials Chemistry and Physics 85 (2004) 113–119
of the FPLAPW methods for the computation of the electronic structure of solids within density functional theory.
In the FPLAPW method, the crystal unit cell is divided into
two parts: atomic spheres centered on the atomic sites and
an interstitial region. Inside the atomic spheres, the basis
117
set used to describe electronic states employs atomic like
functions, while in the interstitial region plane waves are
used. Exchange and correlation effects were treated using
the generalized gradient approximation (GGA) as described
by Perdew and coworkers [18,19].
4.2. Structure optimization
Following the XRD results, Fe3 Si forms a cubic structure
in the Im3m (2 2 9) space group, whereas FeNb crystallizes
in the rhombohedral R-3m (1 6 6) structure. Accordingly, we
first undertake the structural optimization of both materials.
The total energy was calculated self-consistently for different volumes, leading to the energy versus volume plot that
we fit to the Murnaghan equation of state [20] to obtain the
equilibrium volume, total energy and bulk modulus. In all
calculations a set of 200–30,000 points where used in the
irreducible Brillouin zone (IBZ), and the self-consistent cycles where terminated when both the energy and the charge
density reaches an accuracy of 10−5 . In the case of FeNb,
the c/a ratio was first optimized following the same method,
but fitting the energy versus c/a ratio-resulting curves to
a second-order polynomial equation. Fig. 6a and b shows
the total energy versus volume change for cubic Fe3 Si and
rhombohedral FeNb. Table 2 reports the equilibrium calculated lattice parameters, total energies and bulk modulus for
both materials.
Our interest being on the influence of boron impurities in
the Fe3 Si lattice on its electronic properties, first we have
to find the new lattice parameters after the introduction of
boron atoms in the structure. For that we have used the
suppercell method: we reduced the symmetry of the lattice by making equivalent positions nonequivalent, we created a translational reproductions of the unit cell, then we
removed at each time an iron atom and replaced it by a
boron one. Hence, we created doped structures with 8, 16
and 25 at.% of boron. For the Brillouin zone sampling we
followed the rule of thumb that stipulates that the size of
the k-mesh for the suppercell must be that of the unit cell
(original one) multiplied by the ratio of the volumes of simulation cells. This, with the fact that the computational time
grow rapidly with growing k-mesh, we were not able to simulate smaller amounts of boron. As before, we plotted the
energy versus volume and fitted to the Murnaghan equation
of state. The resultant structural properties are presented in
Table 2.
Table 2
Calculated equilibrium volume and bulk modulus for FeNb, Fe3 Si and
B-doped Fe3 Si compounds
Fig. 6. Total energy vs. volume change for (a) cubic Fe3 Si and (b)
rhombohedral FeNb fitted to the Murnaghan equation of state.
Material
Equilibrium volume (Å3 )
Bulk modulus (Gpa)
FeNb
Fe3 Si
Fe3 Si–8 at.% B
Fe3 Si–16 at.% B
Fe3 Si–25 at.% B
28.63
175.32
167.42
160.61
154.81
177.97
183.61
195.21
235.40
252.25
118
M. Cherigui et al. / Materials Chemistry and Physics 85 (2004) 113–119
Fig. 7. Electronic density of state of (a) Fe3 Si and (b) FeNb.
4.3. Electronic structure
Once the crystal structures optimized, we focus on the
electronic properties. We run the self-consistent cycle for
each structure, then the electronic density of state (DOS)
was calculated using the modified tetrahedron method of
Blöchl et al. [21]. Fig. 7 shows the electronic density of
state of both alloys (zero represents the Fermi level). The
DOS histograms show the bonding state orbitals near the
Fermi level and the top of the density curve with the antibonding states. The major difference that we can assert
is the presence in the Fe3 Si histogram of a minimum of
density just beyond the Fermi level, separating the peak of
the antibonding states, whereas for FeNb the Fermi level
is directly followed by a pronounced peak. In both compounds, the antibonding states are the localized d states:
4d-states from iron atoms in both cases, and 3d-states from
niobium atoms in the second. Another difference is that the
peak of the lower energy part of the DOS present for Fe3 Si
and which is due mainly to the tightly bounded s states, is
not visibly separated from the remaining parts of the DOS
for the FeNb alloy. Accordingly, at finite temperatures, the
antibonding states are much more populated in the case
of FeNb alloy. The minimum of density beyond the Fermi
Fig. 8. (a)–(c) Electronic density of state of boron-doped Fe3 Si alloy.
level separating the bonding from the antibonding states in
the iron silicon alloy denotes its structural stability.
The density of state of B-doped Fe3 Si is shown in
Fig. 8a–c. We notice that the shape of DOS histogram is different: the addition of boron atoms annihilate gradually the
minimum present near the Fermi level. In fact, we see that
this minimum is slightly reduced when the boron concentration is of 4 at.%, and is less pronounced for 16 at.% and
disappears for 25 at.% of boron. At the same time, we note
that the peak present in the bonding state is progressively
M. Cherigui et al. / Materials Chemistry and Physics 85 (2004) 113–119
119
Fe3 Si. It has been shown that the introduction of boron
impurities into the alloy matrix leads to the lowering of
the structural stability, and makes its electronic density
of state more comparable to the one corresponding to the
niobium-iron structure.
3. Experimental results have shown clearly that FeNb alloys
can be easily obtained in an amorphous state. According
to the results obtained, and since the crystalline structure
stability is intimately bound to the electronic structure
properties, the addition of boron in the Fe3 Si alloy matrix
will improve its amorphization ability.
References
Fig. 9. Variation of bulk modulus values with boron content.
pronounced, especially, at the Fermi level the density grows
with increasing boron concentration. This is in conformity
with the calculated values of bulk modulus. In fact, this later
increases with the amount of boron, as shown in Fig. 9.
The fact that the antibonding states are more populated
suggests that the addition of boron weaken the crystalline
structure stability, even it reinforce the bonds as stated by
the increasing bulk modulus. Furthermore, the lowest energy
states are progressively connected with the remainder part of
the histogram. Both changes make the electronic density of
state to be similar to the FeNb one. Therefore, the bonding
nature in the Fe3 Si tends to have the same behavior than that
of FeNb owing to the presence of boron.
5. Conclusions
This work is a preliminary contribution to the iron-based
thermally sprayed coatings intended for magnetic applications. The following conclusions can be drawn:
1. A combination of experimental and numerical studies on
deposits obtained by HVOF and APS thermal spraying
of Fe3 Si and FeNb alloys has allowed to compare the
Fe3 Si(B) and FeNb structures. Contrary to the cases of
the microcrystalline Fe3 Si deposits, thermally sprayed
FeNb made it possible to obtain coatings with a partially
amorphous structure.
2. Boron atoms were used to improve the aptitude of the
Fe3 Si alloy to form an amorphous phase. In this perspective, first-principle calculations were elaborated to investigate the electronic structure of crystalline FeNb and
[1] J.C. Perron, Constantes Physico-Chimiques, Techniques de l’Ingénieur, vol. K4, Imprimerie Strasbourgeoise, Schiltigheim, France,
1994.
[2] F.E. Luborsky, Amorphous Metallic Alloys, Butterworths, London,
1983.
[3] R.C. O’Handley, J. Megusar, S.W. Sun, Y. Hara, N.J. Grant, J. Appl.
Phys. 57 (1985) 3563.
[4] A. Borisova, Y. Borisov, V. Korzhyk, V. Bobrik, in: Proceedings of
the ITSC’95, Kobe, May 1995, p. 749.
[5] M. Nakayama, H. Ito, R. Nakamura, M. Toh, A. Ohmori, in:
Proceedings of the ITSC’95, Kobe, May 1995, p. 1063.
[6] H. Herman, S. Sampath, in: Proceedings of the Second Plasma
Technik Symposium, Lucerne, Switzerland, 1991, p. 63.
[7] G. Föoex, J. Phys. Radium 9 (1938) 37.
[8] R. Benoit, J. Chim. Phys. 52 (1955) 119.
[9] J.C. Swartz, R. Kossowsky, J.J. Haugh, R.F. Krause, J. Appl. Phys.
52 (1981) 3324.
[10] T.V. Larionova, O.V. Tolochko, A.S. Zhuravley, Glass Phys. Chem.
21 (1995) 297.
[11] T.V. Larionova, O.V. Tolochko, N.O. Gonchukova, E.V. Novikov,
Glass Phys. Chem. 22 (1996) 248.
[12] N.E. Fenineche, M. Cherigui, A. Kellou, H. Aourag, C. Coddet, in:
Proceedings of the Thermal Spray, ASA, Orlando, 2003, p. 1409.
[13] G.K. Wertheim, V. Jaccarino, J.H. Wernick, J.A. Seitchik, H.J.
Williams, R.C. Sherwood, Phys. Lett. 18 (1965) 89.
[14] L.K. Varga, F. Mazaleyrat, J. Kovac, A. Kàkay, Mater. Sci. Eng. A
304–306 (2001) 946.
[15] M. Abdellaoui, C. Djega-Mariadassou, E. Gaffet, J. Alloys Compd.
259 (1997) 241.
[16] D.D. Koelling, B.N. Harmon, J. Phys. Solid State Phys. 10 (1977)
3107.
[17] P. Blaha, K. Schwarz, G.K.H. Madsen, D. Kvasnicka, J. Luitz,
WIEN2k, An Augmented Plane Wave + Local Orbitals Program
for Calculating Crystal Properties, Vienna University of Technology,
Austria, 2001.
[18] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson,
D.J. Singh, C. Fiolhais, Phys. Rev. B 46 (1992) 6671.
[19] J.P. Perdew, Y. Wang, Phys. Rev. B 45 (1992) 13244.
[20] F.D. Murnaghan, Proc. Nat. Acad. Sci. U.S.A. 30 (1944) 244.
[21] P.E. Blöchl, O. Jepsen, O.K. Andersen, Phys. Rev. B 49 (1994)
16223.